Measurement and Correlation of the Solubility of Sodium 2

Jan 19, 2016 - A synergistic effect on CTSNa solubility was observed with the maximum solubility at solute-free mole fraction of ethanol x 2 0 = 0.208...
0 downloads 0 Views 483KB Size
Article pubs.acs.org/jced

Measurement and Correlation of the Solubility of Sodium 2‑Chlorotoluene-4-sulfonate in Selected Pure Solvents and Aqueous Organic Solutions Shi-Liang Liu, Hai-Ming Li, Guang Han, Yuan-Xiao Li, Guo-Liang Zhang, Feng-Bao Zhang, and Qing Xia* School of Chemical Engineering and Technology, Tianjin University, Tianjin, 300072, P. R. China S Supporting Information *

ABSTRACT: The solubilities of the sodium 2-chlorotoluene4-sulfonate (CTSNa) in water, methanol, formic acid, dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF), binary (methanol + water) mixed solvents, and binary (ethanol + water) mixed solvents were determined experimentally over the temperature range from (280 to 335) K at atmospheric pressure using a dynamic method. The mole fraction of water in the two solvent mixtures ranged from 0.0000 to 1.0000 and 0.2006 to 1.0000, respectively. A synergistic effect on CTSNa solubility was observed with the maximum solubility at solute-free mole fraction of ethanol x02 = 0.2082 and solute-free mole fraction of methanol x03 = 0.3985, respectively. The solubility data were correlated with the thermodynamic electrolyte nonrandom two-liquid (E-NRTL) model and the root-mean-square deviations of solubility temperature varied from (0.24 to 1.10) K.

1. INTRODUCTION Sodium 2-chlorotoluene-4-sulfonate (CTSNa), a white-flaky crystal whose molecular formula is C7H6ClNaO3S and molecular weight is 228.62 g/mol, is used extensively as intermediates for dyes, pigments, pesticides, and medicines.1−6 In general, CTSNa can be prepared by the chlorination of ptoluenesulfonic acid (PTSA) with concentrated sulfuric acid,6,7 and then condensed by NaCl. The product formed by this method ordinarily contains impurities such as sodium 3,5dichloro-4-methylbenzenesulfonate and sodium p-toluenesulfonate. Crystallization which is a relatively low-cost and green process is used in the industry to purify CTSNa. It is necessary to know its solubility in pure water and other common solvents for designing an optimized crystallization process and crystallizer. However, information on the solubility of CTSNa is sparse in the literature. Therefore, it is important to acquire such information. In this work, the solubility of CTSNa in the pure solvents and aqueous organic solutions has been measured using a dynamic method. The solid−liquid-equilibrium (SLE) temperature was investigated ranging from 280 to 335 K at atmospheric pressure. The experimental data were correlated with the electrolyte nonrandom two-liquid (E-NRTL) model8 which has been widely used for representation of phase behavior of aqueous and nonaqueous electrolyte solutions. Meanwhile the E-NRTL model parameters were obtained. The synergistic effect that appeared in this study was also discussed.

Deionized water, the electrolytic conductivity of which is 18 MΩ·cm, was obtained from Nankai Chemical Reagents Co. (Tianjin, China). Detailed information on the materials used in the experiment is listed in Table 1, and all of the solvents were used without further purification. 2.2. Synthesis and Characterization of CTSNa. The processes for the synthesis of CTSNa were as follows. The reaction was carried out in a 1 L borosilicate glass reactor equipped with an electric agitator, a dropping funnel, and a water condenser. The assembly was kept in a constant temperature bath. PTSA and 36% hydrochloric acid were dissolved in the solvent, 35% hydrogen peroxide was added dropwise for 4 h, and the reaction mixture was kept at 318 K. After the end of the reaction, the pH value of the solution was adjusted to about 7, and then NaCl was mixed into the solution. The mixture solution was left to stand for 2 h at room temperature and then filtrated through a Buchner funnel to obtain the crude product. The crude product was further recrystallized several times from water. Finally, a white-flaky crystal of CTSNa was dried in vacuum at 353 K for 12 h. The synthesized CTSNa was characterized by 1H NMR and FTIR spectroscopic techniques. The 1H NMR of synthesized CTSNa was recorded on an AVANCE III NMR spectrometer. The 1H NMR spectrum of synthesized CTSNa is shown in Figure S1 in Supporting Information. 1H NMR (400 MHz, D2O, TMS) δ: 7.72 (s, 1H), 7.53 (d, 1H), 7.37 (d, 1H), 2.34

2. EXPERIMENTAL SECTION 2.1. Materials. The CTSNa was synthesized in our laboratory, and the process is described in section 2.2.

Received: July 29, 2015 Accepted: January 8, 2016

© XXXX American Chemical Society

A

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Source and Mass Fraction Purity of the Materials Used in the Experiments chemical name CTSNa methanol ethanol DMF DMSO formic acid PTSA hydrochloric acid hydrogen peroxide NaCl a

source synthesized in our laboratory Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co., Guangfu Chemical Reagents Co.,

Tianjin, Tianjin, Tianjin, Tianjin, Tianjin, Tianjin, Tianjin, Tianjin, Tianjin,

initial mass fraction purity (%)

purification method

analysis method

99.9 99.8 99.5 99.5 98 99.9 36 35 99.5

recrystallization none none none none none none none none none

HPLCa GCb GCb GCb GCb GCb HPLCa

China China China China China China China China China

HPLC: high performance liquid chromatography. bGC: gas−liquid chromatography.

3.1. Chemical Equilibrium Relationships. With the assumption that all salts completely dissociate in the solution, one can obtain

(s, 3H). IR spectrum was recorded on a Nicolet iS50 FT-IR spectrometer (Thermo Fisher Scientific Inc. USA), and the spectrum was found to be in good agreement with those reported in the National Institute of Advanced Industrial Science and Technology (AIST).9 The purity of the synthesized CTSNa was determined by high performance liquid chromatography (HPLC) analysis (Hitachi L-7100, Japan). A Hypersil BDS C18 column (250 mm·4.6 mm, 5 μm, Thermo) and a UV detector (Hitachi L7400, Japan) at 224 nm length were adopted. HPLC conditions were as follows: water, 0.1% trifluoroacetic acid (v/v) to 20% acetonitrile (flow, 1 mL·min−1). The mobile phase was filtered through a filter (PTFE 0.45 μm) and degassed using ultrasound before passing through the column. The detected mass fraction purity of the synthesized CTSNa was greater than 0.990. After the experiment, a white crystal precipitated from the cooled solvent. The crystal was characterized by 1H NMR. Compared with characterization results before the solubility experiment, the crystal shows no change in the process of dissolution. The 1H NMR spectra for the equilibrium solid phase have been provided in the Supporting Information (Figure S2 and S3). 2.3. Solubility Measurement. The dynamic method10 which was assisted by the laser monitoring observation technique was used for solubility determination at atmospheric pressure.11,12 The operating procedure was performed as the following steps: First, prepared solute and solvent were weighed by an electronic analytical balance (Gibertini, Crystal 200, Italy, the standard uncertainty of 0.0001 g). Then they were slowly mixed and heated in a continuous-stirred jacketed vessel whose volume ranges from 100 to 250 mL according to the various solvent. The temperature of the system was controlled by a refrigerated/heating circulator (Julabo FP45HE, Germany, temperature stability ±0.01 K) and detected by a platinum resistance thermometer (PT-100, the standard uncertainty of 0.01 K). The heating rate would be less than 0.1 K·h−1 near the SLE temperature. A laser monitoring system (JS2-1009016, Beijing, China) was used to observe the course of dissolution and determine the SLE temperature at which the solid just disappeared and the intensity of the laser beam reached a maximum.

M v+X v−(s) ⇄ v+M+(l) + v−X −(l)

(1)

where v denotes the electrolyte stoichiometric coefficient. Mv+Xv− indicates the solid salt consisting of v+ cations M+ and v− anions X− in the crystal. On the basis of eq 1, the solubility product constant Ks can be defined: K s = a+ν+a−ν− = (γ+x+)ν+ (γ−x−)ν−

(2)

where Ks, a, γ, and x denote the solubility product constant, the ion activity, the ion activity coefficient, and the dissociated ion solubility in units of mole fraction, respectively. The subscripts (+) and (−) refer to cation and anion. Furthermore, the solubility product of the solid in the binary solvent mixture which is a function of temperature and solvent mixture composition can be written as follows:13 ln Ksm , n = (xm0 A m + (1 − xm0 )A n) + (xm0 Bm + (1 − xm0 )Bn) /T

(3)

where T is absolute temperature in Kelvin; x0m is the solute-free mole fraction of m component in the binary solvent mixtures. Am, An, Bm, and Bn are constants obtained via solubility data modeling, m and n denote two components of a binary solvent mixture. Particularly, the solubility product in pure solvent m, Ksm, can be expressed as below: ln Ksm = A m + Bm /T

(4)

where Am and Bm are constants obtained by regressing solubility data; m = 1, 2, 3, 4, 5, 6 representing water, methanol, ethanol, formic acid, DMSO, and DMF, respectively. 3.2. The E-NRTL Thermodynamic Model. In this work, the E-NRTL model was used to calculate the activity coefficients in eq 2. The model was originally proposed by Chen14 for aqueous electrolyte systems. It was later extended to mixed solvent electrolyte systems by Mock.15 The model is based on two fundamental assumptions: one for like-ion repulsion assumption and the other for local electroneutrality assumption.16 The excess Gibbs free energy expression contains two contributions: one contribution for the long-range ion−ion interactions (G*m E,PDH) presented by the Pitzer-Debye−Hückel model, and the other for the short-range interaction (G*m E,lc) presented by nonrandom two-liquid (NRTL) theory.16 The E-NRTL model is described by the following expression.

3. THERMODYNAMIC MODELING BASICS In addition to the experimental efforts to measure solubility data of solute in the solvents, suitable thermodynamic equations should be used to correlate the data. In this work, the experimental results were correlated using the E-NRTL model. B

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data Gm*E = Gm*E,PDH + Gm*E.lc

Article

optimization process was aimed at minimizing the following objective function.

(5)

Taking the appropriate derivative form, eq 5 leads to the following equation: ln γi* = ln γi*PDH + ln γi*lc

N

σ = [∑ (T exp − T )2 /(N − 1)]0.5

where σ denotes the root-mean-square deviation between Texp and T. Texp, T, and N represent the experimental equilibrium temperature, calculated equilibrium temperature, and the number of total experimental points, respectively. All the experimental data were used for objective function minimization except for the solubility data of CTSNa in water, since the modeling result for CTSNa in binary solvent mixtures was not satisfactory when taking this set of data into account.

The Pitzer-Debye−Hückel contribution ln γi*PDH is used to represent the long-range interaction contribution and is given by ⎛ 1000 ⎞1/2 ⎡⎛ 2zi2 ⎞ PDH * ⎟ ln(1 + ρIx1/2) ln γi = −⎜ ⎟ A φ ⎢⎜ ⎢ M ρ ⎝ B ⎠ ⎠ ⎣⎝ +

zi2Ix1/2 − 2Ix3/2 ⎤ ⎥ 1 + ρIx1/2 ⎥⎦

4. RESULTS AND DISCUSSION Table 2 presents the mole fraction solubility of CTSNa in the five pure solvents (water, methanol, formic acid, DMSO, and

(7)

where MB, Aφ, zi, ρ and Ix refer to the molecular weight of solvent B, the Pitzer-Debye−Hückel parameter, the charge number of ion i, the closest approach parameter and the ionic strength, respectively.16 The activity coefficient equation for cations is given by16 1 ln γ *lc = zc c

Table 2. Mole Fraction Solubility of CTSNa (x1) at Temperature (Texp) and Pressure 0.1 MPa in Water, Methanol, Formic acid, DMF, DMSOa



Xa ′ ⎞ ∑k XkGkc , a ′ cτkc , a ′ c ⎟⎟ ⎝ ∑a ″ Xa ″ ⎠ ∑k XkGkc , a ′ c

x1·103

Texp/K



4.158 4.366 4.781 5.302 5.922

281.0 284.0 288.1 292.2 296.5

4.095 4.474 4.704 4.976

287.1 293.0 297.1 302.6

36.953 38.312 39.813 41.455

283.0 288.9 295.9 301.4

20.335 20.572 20.841 21.159

280.6 286.3 291.3 297.6

85.990 86.837 87.787

294.0 299.6 304.9

∑ ⎜⎜ a′

+

B′

∑ XG τ ⎞ XBGcB ⎛ ⎜⎜τcB − k k kB kB ⎟⎟ ∑k XkGkB ⎝ ∑k XkGkB ⎠ ⎛

+

Xc ′ ⎞ XaGca , ca ′ ⎟⎟ ⎝ ∑c ″ Xc ″ ⎠ ∑k XkGka , c ′ a

∑ ∑ ⎜⎜ a

c′

⎛ ⎞ ∑ XG τ ⎜⎜τca , c ′ a − k k ka , c ′ a ka , c ′ a ⎟⎟ ∑k XkGka , c ′ a ⎠ ⎝

(8)

The activity coefficient equation for anions is given by16 1 ln γ *lc = za a



Xc ′ ⎞ ∑k XkGka , c ′ aτka , c ′ a ⎟⎟ ⎝ ∑c ″ Xc ″ ⎠ ∑k XkGka , c ′ a

∑ ⎜⎜ c′

+

∑ B

∑ XG τ ⎞ XBGaB ⎛ ⎜⎜τaB − k k kB kB ⎟⎟ ∑k XkGkB ⎝ ∑k XkGkB ⎠ ⎛

+

(11)

i=1

(6)

Xa ′ ⎞ XcGac , a ′ c ⎟⎟ ⎝ ∑a ″ Xa ″ ⎠ ∑k XkGkc , a ′ c

∑ ∑ ⎜⎜ c

a′

⎛ ⎞ ∑ XG τ ⎜⎜τac , a ′ c − k k kc , a ′ c kc , a ′ c ⎟⎟ ∑k XkGkc , a ′ c ⎠ ⎝

(9)

Texp/K

water 6.607 301.2 7.334 305.6 8.352 310.7 9.197 314.6 10.144 318.7 methanol 5.240 306.1 5.498 311.8 5.785 316.7 6.066 321.5 formic acid 43.069 307.3 44.622 313.0 45.785 319.5 47.469 324.9 DMF 21.578 304.8 22.285 310.7 22.850 314.8 23.457 319.0 DMSO 88.737 310.6 89.424 314.5 90.228 319.2

x1·103

Texp/K

11.147 12.198 13.862 15.374

322.4 326.2 331.1 335.1

6.369 6.696

327.0 333.1

49.098 50.313

331.1 335.5

24.050 25.210 26.849

322.1 327.6 333.3

91.162 92.892 94.676

323.5 330.2 335.2

a

The standard uncertainty of the measurement temperature is u(T) = 0.3 K (0.68 level of confidence).The standard uncertainty of the measurement pressure u(p) = 0.002 MPa (0.68 level of confidence). The expanded uncertainty of the solubility measurement is U(x) = 10−5 (0.95 level of confidence).

where Gji,ki = e and the nonrandomness factor α and the energy interaction parameter τij are the E-NRTL model parameters which are expressed as a function of temperature. The format of τij is written as follows:16 (−αji,kiτji,ki)

τij = aij + bij /T

x1·103

(10)

where aij and bij are constants representing the temperature dependence of τij. Generally, the values of the nonrandomness factors for solvent−solvent and water−salt were fixed to 0.3 and 0.2 in the theoretical treatment, respectively. 3.3. Correlation of Experimental Data. The Nelder− Mead Simplex Method17 combined with Matlab (Mathwork, MA) is used to determine the E-NRTL model parameters. The

DMF) from (280 to 335) K at atmospheric pressure. The relationship between the solubility and temperature are plotted in Figure 1. From Table 2 and Figure 1, it can be found that solubilities of CTSNa increase with temperature and the solubilities in the formic acid and water show the strongest positive dependency on temperature. Thus, formic acid and water are favorable for crystallization. C

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. Plot of mole fraction solubility of CTSNa (x1) vs temperature (T) at different solute-free ethanol mole fraction (x02) of aqueous solution: points, experimental values, (△) x02 = 0.0000; (+) x02 = 0.1002; (□) x02 = 0.1985; (○) x02 = 0.3986; (▽) x02 = 0.5975; (☆) x02 = 0.7994; , calculated from the E-NRTL model.

Figure 1. Plot of mole fraction solubility of CTSNa (x1) vs temperature (T) in pure solvents: (□) DMSO; (○) formic acid; (△) DMF; (◊) water; (▽) methanol; , calculated from the ENRTL model.

solubility increases with enhance of temperature. Because CTSNa almost does not dissolve in the pure ethanol, its solubility cannot be precisely measured and the ethanol whose mole fractions arrange from 0 to 0.7994 are chosen. The solubility of CTSNa increases as augment of x02 and reaches the maximum value at x02 = 0.2082, and then it decreases with a further increase of x02. As for the solubility in methanol−water solvents, the experimental data are shown in Table 4 and Figure 3, where

With respect to the solubility of CTSNa in ethanol−water mixed solvent, its results are illustrated in Table 3 and Figure 2, where x02 is the solute-free mole fraction of ethanol in the binary (ethanol + water) mixed solvents. As shown in Figure 2, the Table 3. Mole Fraction Solubility of CTSNa (x1) at Temperature (Texp) and Pressure 0.1 MPa in Different Solute-free Ethanol Mole Fraction (x02) of Aqueous Solutiona x1·103

Texp/K

x1·103

Texp/K

x1·103

Texp/K

16.312 18.209

328.9 333.8

Table 4. Mole Fraction Solubility of CTSNa (x1) at Temperature (Texp) and Pressure 0.1 MPa in Different Solute-free Methanol Mole Fraction (x03) of Aqueous Solutiona

x02

4.722 5.698 6.700 8.053

284.0 291.2 296.2 302.3

5.707 6.311 7.416 9.218

284.6 287.3 291.4 298.0

5.529 6.610 7.522 8.436

286.8 291.3 295.4 299.3

3.052 3.820 4.460 5.128

287.3 293.1 298.4 303.1

1.264 1.553 1.720 1.898

286.9 296.3 299.7 302.9

= 0.1002 9.372 308.4 10.964 313.7 12.887 319.2 14.608 324.3 x02 = 0.1985 10.896 302.9 12.617 307.3 14.487 312.0 16.735 316.7 x02 = 0.3986 9.439 303.0 10.644 307.4 12.089 312.3 13.751 317.3 x02 = 0.5975 5.828 307.5 6.642 312.1 7.539 316.5 8.551 321.2 x02 = 0.7994 2.124 307.0 2.361 310.4 2.633 314.3 2.936 318.2

x1·103

Texp/K

19.166 21.839 24.898

321.9 326.8 332.1

4.151 4.840 5.734 6.557

282.4 287.1 293.1 297.3

15.720 17.705 20.025

322.7 327.7 333.1

4.257 4.641 5.027 5.824

279.9 282.5 285.0 290.1

9.585 10.757 12.024

325.9 329.8 334.6

4.053 4.658 5.281 5.913 6.589

283.0 287.6 291.7 295.1 299.2

3.272 3.659 4.098 4.485

322.4 326.6 331.4 334.8

4.109 4.318 5.289

282.1 285.5 293.5

a

x1·103

Texp/K

x03

0.2082 302.3 306.9 311.4 316.4 0.3985 293.9 298.6 303.0 308.2 0.6023 303.4 307.2 311.6 314.8 318.5 0.8011 300.0 307.1 313.0

= 7.513 8.559 9.765 11.130 x03 = 6.644 7.545 8.610 9.940 x03 = 7.299 8.032 8.881 9.597 10.501 x03 = 6.060 6.940 7.681

x1·103

Texp/K

12.645 14.410 16.243

321.2 326.1 330.2

11.525 13.171 14.976

313.8 319.0 323.9

12.210 13.871 15.393

324.7 330.1 334.3

8.490 9.321 10.147

318.0 324.7 330.1

a

The standard uncertainty of the measurement temperature is u(T) = 0.03 K (0.68 level of confidence). The standard uncertainty of the measurement pressure u(p) = 0.002 MPa (0.68 level of confidence). The expanded uncertainty of the solubility measurement is U(x) = 10−5 (0.95 level of confidence).

The standard uncertainty of the measurement temperature is u(T) = 0.03 K (0.68 level of confidence).The standard uncertainty of the measurement pressure u(p) = 0.002 MPa (0.68 level of confidence). The expanded uncertainty of the solubility measurement is U(x) = 10−5 (0.95 level of confidence). D

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 3. Plot of mole fraction solubility of CTSNa (x1) vs temperature (T) at different solute-free methanol mole fraction (x03) of aqueous solution: points, experimental values, (+) x03 = 0.0000; (□) x03 = 0.2082; (○) x03 = 0.3985; (△) x03 = 0.6023; (▽) x03 = 0.8011; (◊) x03 = 1.0000; , calculated from the E-NRTL model.

Figure 5. Solubility of CTSNa (x1) in binary solvent mixtures (methanol + water) at three temperatures: (△) T = 317.15 K; (○) T = 307.15 K; (□) T = 297.15 K; x03, solute-free methanol mole fraction of CTSNa in the binary solvent mixture. , calculated from the ENRTL model.

x03 is the solute-free mole fraction of methanol in the binary (methanol + water) mixed solvents. Analogous to the (ethanol + water) system, the solubility of CTSNa in methanol enhances as temperature is increasing, but the difference of solubility for each condition is relatively small when the temperature is less than 295 K. Furthermore, the solubility enhances first and decreases subsequently with x03; its maximum value is reached when x03 = 0.3985. According to the above experimental results, there is enhanced solubility in (methanol + water) and (ethanol + water) systems, known as the synergistic effect,18 which means that the power of dissolution of mixed solvent is much higher than that of their individual components at a particular composition. Plot of calculated solute solubility, x1 versus solvent composition reveals the positive synergistic effect in (methanol + water) and (ethanol + water) systems as presented in Figure 4 and Figure 5, respectively. Many experiments indicate that the synergistic effect exists in the systems

containing alcohol.19−21 Alcohol and water were associated liquids which had different association structure groups.22,23 They might be expected to form new association structures after mixing. Also, each association structure model changed with the water content of alcohol−water mixtures and temperature.24−27 Ethanol, methanol, and water used in our work are associated liquids whose molecules are held together by hydrogen bonds which can interact with each other. Probably the hydrogen bonds competition between like and unlike molecules is an important factor to determine the phase behavior.28,29 Table 5 contains the parameters of the solubility product for CTSNa in different solvents, which is defined by eq 4. Table 6 Table 5. Parameters of Solubility Products for CTSNa in Different Pure Solvents and Aqueous Organic Solutions (Defined by Equation 4) solvent system

Am

Bm/K

water (m = 1) ethanol (m = 2) methanol (m = 3) formic acid (m = 4) DMSO (m = 5) DMF (m = 6)

−18.926 298.24 −145.11 2.4034 493.55 60.035

−2710 −47460 92004 −1639.3 −249438 −17282

represents the correlated E-NRTL model parameters. The rootmean-square deviations of SLE temperature described by the ENRTL model are summarized in Table 7. The root-meansquare deviations of SLE temperature vary from (0.24 to 1.10) K, indicating the goodness of fit of the solubility curves by means of the E-NRTL model.

5. CONCLUSIONS Solubility data of CTSNa in the pure solvents and binary alcohols + water solvent mixtures from 280 to 335 K have been determined and discussed. Solubility increased with temperature. The solvents had the solubility order DMSO > formic acid > DMF > water > methanol, with CTSNa being nearly insoluble in ethanol. The synergistic effect on CTSNa solubility

Figure 4. Solubility of CTSNa (x1) in binary solvent mixtures (ethanol + water) at three temperatures: (△) T = 327.15 K; (○) T = 307.15 K; (□) T = 287.15 K; x02, solute-free ethanol mole fraction of CTSNa in the binary solvent mixture. , calculated from the E-NRTL model. E

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 6. E-NRTL Model Parameters for CTSNa in Different Pure Solvents and Aqueous Organic Solutions (Defined by Equation 7) i

j

aij

aji

bij

bji

aij = aji

CTSNa CTSNa water CTSNa water CTSNa CTSNa CTSNa

water ethanol ethanol methanol methanol formic acid DMSO DMF

1.3333 16.842 129.67 −5.2644 −26.224 1155.9 1484.6 20943

5.5670 156.54 5.4832 −95.600 10.959 3.2011 230.36 35.454

−2792.3 4144.1 −27693 3972.6 −10257 6807.1 975984 −5137.2

7033.1 −49185 −1971.0 55986 −525.79 −405.34 −121788 −9024.7

0.2 −0.033691 0.3 0.028885 0.3 0.010733 0.007135 0.046413

Table 7. Root-Mean-Square Deviations σ from the Description by the E-NRTL Model solvent system water water + water + water + water + water + DMSO

0.1002 0.1985 0.3986 0.5975 0.7994

ethanol ethanol ethanol ethanol ethanol

σ/K 0.39 1.07 1.10 0.87 0.84 0.24 0.39

solvent system water + 0.2082 water + 0.3985 water + 0.6023 water + 0.8011 methanol formic acid DMF

methanol methanol methanol methanol

(6) Qiu, W. Z.; Zou, G.; Wang, X. H.; Xu, Q. Z.; Chen, Z. W.; Liu, R.; Zhao, W. Method for Recycling and Reusing Two Main Byproducts in Production of 2-Amino-5-Chloro-p-toluene sulfonic Acid [P]. CN Patent 101935294, 2011. (7) Tonne, P.; Mueller, E. R.; Goerth, H. Preparation of 1-Alkyl-2Chloro-5-nitrobenze -ne-4-sulfonic Acids [P]. U.S. Patent 4427600, 1984. (8) Chen, C. C.; Evans, L. B. A Local Composition Model for the Excess Gibbs Energy of Aqueous Electrolyte Systems. AIChE J. 1986, 32, 444−459. (9) SDBSWeb: http://sdbs.db.aist.go.jp (accessed January 5, 2015). (10) Ma, P. S.; Xia, Q. Determination and Correlation for Solubility of Aromatic Acids in Solvents. Chin. J. Chem. Eng. 2001, 9, 39−44. (11) Xia, Q.; Zhang, F. B.; Zhang, G. L.; Ma, J. C.; Zhao, L. Solubility of Sebacic Acid in Binary Water + Ethanol Solvent Mixtures. J. Chem. Eng. Data 2008, 53, 838−840. (12) Xia, Q.; Chen, S. N.; Chen, Y. S.; Zhang, M. S.; Zhang, F. B.; Zhang, G. L. Solubility of Decanedioic Acid in Binary Solvent Mixtures. Fluid Phase Equilib. 2011, 304, 105−109. (13) Wang, L. X.; Xia, Q.; Kang, J.; Du, M. X.; Zhang, G. L.; Zhang, F. B. Measurement and Correlation of Solubilities of Potassium Chloride and Potassium Sulfate in Aqueous Glycerol Solutions. J. Chem. Eng. Data 2011, 56, 3813−3817. (14) Chen, C. C.; Britt, H. I.; Boston, J. F.; Evans, L. B. Local Composition Model for Excess Gibbs Energy of Electrolyte Systems. AIChE J. 1982, 28, 588−596. (15) Mock, B.; Evans, L. B.; Chen, C. C. Thermodynamic Representation of Phase Equilibria of Mixed-Solvent Electrolyte Systems. AIChE J. 1986, 32, 1655−1664. (16) Aspen Physical Property System 11.1; Aspen Technology: Cambridge, 2001. (17) Nelder, J. A.; Mead, R. A. A Simplex Method for Function Minimization. J. Comput. 1965, 7, 308−313. (18) Hildebrand, I. H.; Prausnitz, J. M.; Scott, R. L. Regular and Related Solutions; Van Nostrand Reinhold: New York, 1970. (19) Chen, S. N.; Xia, Q.; Lu, L. F.; Li, D.; Zhang, F. B.; Zhang, G. L. Solubility of Sebacic Acid in Binary Mixtures Containing Alcohols. J. Chem. Eng. Data 2009, 54, 776−780. (20) Chen, S. N.; Xia, Q.; Lu, L. F.; Zhang, M. S.; Chen, Y. S.; Zhang, F. B.; Zhang, G. L. Measurement and Correlation of Solubilities of Decanedioic Acid in C4-C6. J. Chem. Eng. Data 2010, 55, 1411−1415. (21) Domanska, U. Solubility of 2,5-Dimethylphenol and 3,4Dimethylphenol in Binary Solvent Mixtures Containing Alcohols. Fluid Phase Equilib. 1988, 40, 259−277. (22) Franks, F.; Ives, D. J. G. The Structural Properties of AlcoholWater Mixtures. Q. Rev., Chem. Soc. 1966, 20, 1−44. (23) Pratt, L. R.; Laviolette, R. A. Quasi-chemical theories of associated liquids. Mol. Phys. 1998, 94, 909−915. (24) Coccia, A.; Indovina, P. L.; Podo, F.; Viti, V. PMR studies on the structures of water-ethyl alcohol mixtures. Chem. Phys. 1975, 7, 30−40. (25) Monakhova, Y. B.; Mushtakova, S. P.; Kuballa, T.; Lachenmeier, D. W. Investigation into the structural composition of hydroalcoholic solutions as basis for the development of multiple suppression pulse sequences for NMR measurement of alcoholic beverages. Magn. Reson. Chem. 2014, 52, 755−759.

σ/K 0.56 0.70 0.61 0.41 0.89 0.34 0.40

was observed in (CTSNa + ethanol + water) and (CTSNa + methanol + water) systems with maximum solubility at x02 = 0.2082 and x03 = 0.3985, respectively. Also, the synergistic effect may be a result of the hydrogen bonds. The E-NRTL thermodynamic model was used to describe the phase equilibrium of crystal CTSNa in the abovementioned solutions. The calculated solubilities with model parameters were satisfactory, and the root-mean-square deviations of SLE temperature ranged from (0.24 to 1.10) K.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.5b00655. 1 H NMR spectra of synthesized CTSNa before and after dissolution (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +86-22-27400292. Fax: +8622-27408778. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Ralf, W.; Sven, R.; Peter, D.; Alexander, H.; Walter, E.; Birgitt, S.; Gudrun, R. Preparation of Sulfamoyl Steroid Prodrugs with Androgenic Action [P]. U.S. Patent 0282793, 2005. (2) Ioan, P.; Iuliu, C.; Lazar, T.; Ludovic, F.; Tereza Rodica, C. oChlorotoluene [P]. RO Patent 70516, 1981. (3) Guo, S.; Yang, F.; Zhang, L. Synthesis of Herbicide Sulcotrione. Nongyao 2001, 40, 20−21. (4) Rudolf, H.; Juergen, D.; Peter, F.; Walter, H.; Thomas, K.; Ulrich, M.; Matthias, M. G.; Martin, B.; Stanislav, K. Preparation of Sulfonylbenzylimidazoles as Angiotensin II Antagonists [P]. U.S. Patent 5318980, 1994. (5) Nippon Kayaku Co., Ltd. Toluenesulfonic Acid Derivative [P]. JP Patent 60016965, 1985. F

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(26) Ahmed, M. K.; Ali, S.; Wojcik, E. The C-O Stretching Infrared Band as a Probe of Hydrogen Bonding in Ethanol-Water and Methanol-Water Mixtures. Spectrosc. Lett. 2012, 45, 420−423. (27) Weirberg, L.; Zimmormann, J. R. Concentration Dependence of Chemical Exchange and NMR Multiplet Sturcture in Water-Ethanol Association Mixtures. J. Chem. Phys. 1955, 23, 748. (28) Zou, W.; Xia, Q.; Zhao, W.; Zhang, F. B.; Zhang, G. L. Solubility of disodium 4,4′- dinitrostilbene-2,2′disulfonate and sodium 4nitrotoluene-2-sulfonate in aqueous organic solutions and its application feasibility in oxidation stage of DSD acid synthesis. J. Chem. Thermodyn. 2014, 70, 239−244. (29) Ballal, D.; Chapman, W. G. Competition between Intra- and Intermolecular Association of Chain Molecules with Water-like Solvent. J. Phys. Chem. B 2015, 119, 6792−6802.

G

DOI: 10.1021/acs.jced.5b00655 J. Chem. Eng. Data XXXX, XXX, XXX−XXX